Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

LAG3 associates with TCR–CD3 complexes and suppresses signaling by driving co-receptor–Lck dissociation

Subjects

Abstract

LAG3 is an inhibitory receptor that is highly expressed on exhausted T cells. Although LAG3-targeting immunotherapeutics are currently in clinical trials, how LAG3 inhibits T cell function remains unclear. Here, we show that LAG3 moved to the immunological synapse and associated with the T cell receptor (TCR)-CD3 complex in CD4+ and CD8+ T cells, in the absence of binding to major histocompatibility complex class II—its canonical ligand. Mechanistically, a phylogenetically conserved, acidic, tandem glutamic acid–proline repeat in the LAG3 cytoplasmic tail lowered the pH at the immune synapse and caused dissociation of the tyrosine kinase Lck from the CD4 or CD8 co-receptor, which resulted in a loss of co-receptor–TCR signaling and limited T cell activation. These observations indicated that LAG3 functioned as a signal disruptor in a major histocompatibility complex class II-independent manner, and provide insight into the mechanism of action of LAG3-targeting immunotherapies.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: LAG3 inhibits T cell function in the absence of MHC class II ligation.
Fig. 2: LAG3 associates with the TCR–CD3 complex.
Fig. 3: LAG3 association with the TCR–CD3 complex increases LAG3 proximity to the coreceptors.
Fig. 4: LAG3 association with the TCR–CD3 complex increases LAG3 proximity to the coreceptors.
Fig. 5: The ‘EP’ motif in the CT of LAG3 disrupts co-receptor–Lck complex.
Fig. 6: LAG3 dissociates Lck from CD4 by lowering the local pH at the IS.

Similar content being viewed by others

Data availability

All data generated or analyzed during this study are included in this published article and its supplementary information files. Source data are provided with this paper.

References

  1. Andrews, L. P., Marciscano, A. E., Drake, C. G. & Vignali, D. A. LAG3 (CD223) as a cancer immunotherapy target. Immunol. Rev. 276, 80–96 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Maruhashi, T., Sugiura, D., Okazaki, I. M. & Okazaki, T. LAG-3: from molecular functions to clinical applications. J. Immunother. Cancer 8, e001014 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Ruffo, E., Wu, R. C., Bruno, T. C., Workman, C. J. & Vignali, D. A. A. Lymphocyte-activation gene 3 (LAG3): the next immune checkpoint receptor. Semin. Immunol. 42, 101305 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Huard, B. et al. Characterization of the major histocompatibility complex class II binding site on LAG-3 protein. Proc. Natl Acad. Sci. USA 94, 5744–5749 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Workman, C. J. & Vignali, D. A. Negative regulation of T cell homeostasis by lymphocyte activation gene-3 (CD223). J. Immunol. 174, 688–695 (2005).

    Article  CAS  PubMed  Google Scholar 

  6. Workman, C. J. & Vignali, D. A. The CD4-related molecule, LAG-3 (CD223), regulates the expansion of activated T cells. Eur. J. Immunol. 33, 970–979 (2003).

    Article  CAS  PubMed  Google Scholar 

  7. Woo, S. R. et al. Immune inhibitory molecules LAG-3 and PD-1 synergistically regulate T-cell function to promote tumoral immune escape. Cancer Res. 72, 917–927 (2012).

    Article  CAS  PubMed  Google Scholar 

  8. Blackburn, S. D. et al. Coregulation of CD8+ T cell exhaustion by multiple inhibitory receptors during chronic viral infection. Nat. Immunol. 10, 29–37 (2009).

    Article  CAS  PubMed  Google Scholar 

  9. Wherry, E. J. T cell exhaustion. Nat. Immunol. 12, 492–499 (2011).

    Article  CAS  PubMed  Google Scholar 

  10. Workman, C. J., Dugger, K. J. & Vignali, D. A. Cutting edge: molecular analysis of the negative regulatory function of lymphocyte activation gene-3. J. Immunol. 169, 5392–5395 (2002).

    Article  CAS  PubMed  Google Scholar 

  11. Maeda, T. K., Sugiura, D., Okazaki, I. M., Maruhashi, T. & Okazaki, T. Atypical motifs in the cytoplasmic region of the inhibitory immune co-receptor LAG-3 inhibit T cell activation. J. Biol. Chem. 294, 6017–6026 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Baixeras, E. et al. Characterization of the lymphocyte activation gene 3-encoded protein. A new ligand for human leukocyte antigen class II antigens. J. Exp. Med. 176, 327–337 (1992).

    Article  CAS  PubMed  Google Scholar 

  13. Grebinoski, S. & Vignali, D. A. Inhibitory receptor agonists: the future of autoimmune disease therapeutics? Curr. Opin. Immunol. 67, 1–9 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Szent-Gyorgyi, C. et al. Fluorogen-activating single-chain antibodies for imaging cell surface proteins. Nat. Biotechnol. 26, 235–240 (2008).

    Article  CAS  PubMed  Google Scholar 

  15. Perkins, L. A. et al. High-content surface and total expression siRNA kinase library screen with VX-809 treatment reveals kinase targets that enhance F508del-CFTR rescue. Mol. Pharm. 15, 759–767 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Samir, P. et al. DDX3X acts as a live-or-die checkpoint in stressed cells by regulating NLRP3 inflammasome. Nature 573, 590–594 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Fu, G. et al. Metabolic control of TFH cells and humoral immunity by phosphatidylethanolamine. Nature 595, 724–729 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Chen, F., Tillberg, P. W. & Boyden, E. S. Optical imaging. Expansion microscopy. Science 347, 543–548 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Wassie, A. T., Zhao, Y. & Boyden, E. S. Expansion microscopy: principles and uses in biological research. Nat. Methods 16, 33–41 (2019).

    Article  CAS  PubMed  Google Scholar 

  20. Huang, W. Y. C., Ditlev, J. A., Chiang, H. K., Rosen, M. K. & Groves, J. T. Allosteric modulation of Grb2 recruitment to the intrinsically disordered scaffold protein, LAT, by remote site phosphorylation. J. Am. Chem. Soc. 139, 18009–18015 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Bettini, M. et al. Cutting edge: accelerated autoimmune diabetes in the absence of LAG-3. J. Immunol. 187, 3493–3498 (2011).

    Article  CAS  PubMed  Google Scholar 

  22. Huang, C. T. et al. Role of LAG-3 in regulatory T cells. Immunity 21, 503–513 (2004).

    Article  CAS  PubMed  Google Scholar 

  23. Woo, S. R. et al. Differential subcellular localization of the regulatory T-cell protein LAG-3 and the coreceptor CD4. Eur. J. Immunol. 40, 1768–1777 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Workman, C. J. et al. Lymphocyte activation gene-3 (CD223) regulates the size of the expanding T cell population following antigen activation in vivo. J. Immunol. 172, 5450–5455 (2004).

    Article  CAS  PubMed  Google Scholar 

  25. Grosso, J. F. et al. Functionally distinct LAG-3 and PD-1 subsets on activated and chronically stimulated CD8 T cells. J. Immunol. 182, 6659–6669 (2009).

    Article  CAS  PubMed  Google Scholar 

  26. Grosso, J. F. et al. LAG-3 regulates CD8+ T cell accumulation and effector function in murine self- and tumor-tolerance systems. J. Clin. Invest. 117, 3383–3392 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Veillette, A., Bookman, M. A., Horak, E. M. & Bolen, J. B. The CD4 and CD8 T cell surface antigens are associated with the internal membrane tyrosine-protein kinase Lck. Cell 55, 301–308 (1988).

    Article  CAS  PubMed  Google Scholar 

  28. Rudd, C. E., Trevillyan, J. M., Dasgupta, J. D., Wong, L. L. & Schlossman, S. F. The CD4 receptor is complexed in detergent lysates to a protein-tyrosine kinase (pp58) from human T lymphocytes. Proc. Natl Acad. Sci. USA 85, 5190–5194 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Horkova, V. et al. Dynamics of the coreceptor-LCK interactions during T cell development shape the self-reactivity of peripheral CD4 and CD8 T cells. Cell Rep. 30, 1504–1514 e1507 (2020).

    Article  CAS  PubMed  Google Scholar 

  30. Kim, P. W., Sun, Z. Y., Blacklow, S. C., Wagner, G. & Eck, M. J. A zinc clasp structure tethers Lck to T cell coreceptors CD4 and CD8. Science 301, 1725–1728 (2003).

    Article  CAS  PubMed  Google Scholar 

  31. Alberts, I. L., Nadassy, K. & Wodak, S. J. Analysis of zinc binding sites in protein crystal structures. Protein Sci. 7, 1700–1716 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Rigo, A. et al. Interaction of copper with cysteine: stability of cuprous complexes and catalytic role of cupric ions in anaerobic thiol oxidation. J. Inorg. Biochem. 98, 1495–1501 (2004).

    Article  CAS  PubMed  Google Scholar 

  33. Lee, K. H. et al. C9orf72 dipeptide repeats impair the assembly, dynamics, and function of membrane-less organelles. Cell 167, 774–788 e717 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Mitrea, D. M. & Kriwacki, R. W. Phase separation in biology; functional organization of a higher order. Cell Commun. Signal 14, 1 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Storch, S., Pohl, S. & Braulke, T. A dileucine motif and a cluster of acidic amino acids in the second cytoplasmic domain of the batten disease-related CLN3 protein are required for efficient lysosomal targeting. J. Biol. Chem. 279, 53625–53634 (2004).

    Article  CAS  PubMed  Google Scholar 

  36. Johnson, A. O., Lampson, M. A. & McGraw, T. E. A di-leucine sequence and a cluster of acidic amino acids are required for dynamic retention in the endosomal recycling compartment of fibroblasts. Mol. Biol. Cell 12, 367–381 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Uversky, V. N. The alphabet of intrinsic disorder: II. Various roles of glutamic acid in ordered and intrinsically disordered proteins. Intrinsically Disord. Proteins 1, e24684 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Miyazaki, T., Dierich, A., Benoist, C. & Mathis, D. Independent modes of natural killing distinguished in mice lacking Lag3. Science 272, 405–408 (1996).

    Article  CAS  PubMed  Google Scholar 

  39. Kaye, J. et al. Selective development of CD4+ T cells in transgenic mice expressing a class II MHC-restricted antigen receptor. Nature 341, 746–749 (1989).

    Article  CAS  PubMed  Google Scholar 

  40. Singh, S. K. et al. Mapping the interaction between the cytoplasmic domains of HIV-1 viral protein U and human CD4 with NMR spectroscopy. FEBS J. 279, 3705–3714 (2012).

    Article  CAS  PubMed  Google Scholar 

  41. Huppa, J. B. et al. TCR-peptide-MHC interactions in situ show accelerated kinetics and increased affinity. Nature 463, 963–967 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Kaizuka, Y., Douglass, A. D., Varma, R., Dustin, M. L. & Vale, R. D. Mechanisms for segregating T cell receptor and adhesion molecules during immunological synapse formation in Jurkat T cells. Proc. Natl Acad. Sci. USA 104, 20296–20301 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Liedmann, S. et al. Viral suppressors of the RIG-I-mediated interferon response are pre-packaged in influenza virions. Nat. Commun. 5, 5645 (2014).

    Article  CAS  PubMed  Google Scholar 

  44. Bates, M., Huang, B., Dempsey, G. T. & Zhuang, X. Multicolor super-resolution imaging with photo-switchable fluorescent probes. Science 317, 1749–1753 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Keller, R. The Computer Aided Resonance Assignment 1st edn (CANTINA, 2004).

  46. Santoro, M. M. & Bolen, D. W. Unfolding free energy changes determined by the linear extrapolation method. 1. Unfolding of phenylmethanesulfonyl alpha-chymotrypsin using different denaturants. Biochemistry 27, 8063–8068 (1988).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We wish to thank everyone in the Vignali laboratory (Vignali-lab.com; @Vignali_Lab) for all their constructive comments and advice during this project. The authors would like to thank G. Lennon, R. Cross and P. Ingle of the St. Jude Immunology Flow Lab for cell sorting and valuable assistance; A. Yates, D. Falkner and H. Shen from the Immunology Flow Core at the University of Pittsburgh for cell sorting; A. McKenna and K. Forbes for maintenance, breeding and genotyping of mouse colonies at St. Jude Children’s Research Hospital; E. Brunazzi for maintenance, breeding and genotyping of mouse colonies at the University of Pittsburgh; and the staffs of the Shared Animal Resource Center and the Division of Laboratory Animal Services for the animal husbandry A. Philips for assistance with NMR data analysis. Images were acquired, in part, at the Cell and Tissue Imaging Center, and peptide synthesis was performed by the Hartwell Center for Macromolecular Synthesis, which are supported by SJCRH and NCI P30 CA021765. This work was supported by the National Institutes of Health (P01 AI108545, R01 AI129893 to D.A.A.V.; R01 AI144422 to D.A.A.V and C.J.W.), NCI Comprehensive Cancer Center Support CORE grant (CA047904, to D.A.A.V. and R.K.), and ALSAC (to D.A.A.V. and R.K.).

Author information

Authors and Affiliations

Authors

Contributions

D.A.A.V., C.G. and C.J.W. conceptualized the study. C.G., C.J.W., D.M.M., R.K., S.C.W. and D.A.A.V. developed the methodology. C.G., C.J.W., X.L. and H.Z. performed formal analysis. Investigations were performed by C.G., C.J.W., P.-C.C., D.M.M. and K.V. Resources were provided by D.A.A.V., M.P.B., R.K., S.C.W. and J.T. C.G., C.J.W. and D.A.A.V prepared the original draft of the manuscript. All authors wrote, reviewed and edited the final paper. D.A.A.V., R.K. and S.C.W. supervised the project; D.A.A.V. acquired funding.

Corresponding authors

Correspondence to Creg J. Workman or Dario A. A. Vignali.

Ethics declarations

Competing interests

D.M.M. is an employee and stock holder of Dewpoint Therapeutics. D.A.A.V. and C.J.W. have patents covering LAG3, with others pending, and are entitled to a share in net income generated from licensing of these patent rights for commercial development. D.A.A.V. is a cofounder and stock holder at Novasenta, Potenza, Tizona, Trishula; is a stock holder at Oncorus, Werewolf, Apeximmune; has patents licensed and royalties at Astellas, BMS, Novasenta; is a scientific advisory board member at Tizona, Werewolf, F-Star, Bicara, Apeximmune, T7/Imreg Bio; is a consultant at Astellas, BMS, Almirall, Incyte, G1 Therapeutics; has received research funding at BMS, Astellas and Novasenta. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Immunology thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editor: Ioana Visan, in collaboration with the Nature Immunology team.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1

(a) Thymidine incorporation proliferation assay with purified CD4+ T cells from spleen and lymph nodes of Lag3+/+ and Lag3–/– mice stimulated with plate-bound CD3ε and CD28 Abs for 72 h in the absence (a) or presence of isotype control or anti-LAG3 blocking antibodies (b). (c) LAG3 expression of Lag3+/+ CD4+ T cells, isolated as above following activation with CD3ε and CD28 Abs at 48 hr. Cells were gated on a live lymphocyte gate, and CD4. (d) Diagram depicting TIRF microscopy analysis on stimulating lipid bilayer. (e) Western blot analysis and quantification of TCR–proximal signaling events in Lag3+/+ or Lag3–/– CD4+ T cells stimulated for 48 hr with CD3ε and CD28 Abs to induce LAG3 expression, allowed to rest in the presence of IL-2 and then restimulated for 5, 15 and 30 min with anti-CD3. Quantitation determined by percent density of Lag3–/– compared to Lag3+/+. Data in (a) represents the mean of n = 16 (Lag3+/+) and n = 16 (Lag3–/–) individual mice. Data in (b) represents the mean of n = 6 (Lag3+/+) and n = 6 (Lag3–/–) individual mice. Data in (e) represents the mean of 4-8 individual experiments. Statistical analysis performed using Student’s unpaired two-sided t test with P values noted in figures.

Source data

Extended Data Fig. 2

(a) Diagram depicting TIRF microscopy analysis of Lag3–/– CD4+ T cells containing FAP-tagged LAG3 in the presences of MGnBu to visualize the FAP, stimulated on a planar lipid bilayer containing ICAM and biotinylated TCRβ Ab, with streptavidin Alexa 488 to allow for visualization of TCR clustering. (b) Real-time fluorescent TIRFM visualizing LAG3 (red) and TCR (green) of CD8+ T cells, isolated from spleen and lymph nodes of Lag3–/–mice, containing FAP-tagged LAG3 stimulated on a planar lipid bilayer containing TCRβ Ab for 15 minutes (Scale bar = 5 µm). (c) Activated T cells were immunolabeled for Tubulin (green) and mitochondria (Red) and imaged using 3-dimensional super-resolution confocal either pre- (upper) or post-expansion microscopy (lower) and their maximum cell width determined as a measure of fold expansion. (d) Lag3+/+ and Lag3–/– CD4+ and CD8+ T cells, isolated as above, were immunolabeled to detect TCR and LAG3, and subjected to expansion microscopy with the degree of colocalization presented (Scale bar = 10 µm). (e) Coimmunoprecipitation of the TCR–CD3 complex with LAG3 in resting Lag3+/+ and Lag3–/– transgenic CD4+ T cells and stimulated in the presence of peptide and MHC class II. (f) Molecular distribution of TCR and LAG3 in non-stimulated Lag3+/+ CD4+ and CD8+ T cells, isolated as above, as determined by STORM with quantification shown on the right (Scale bar = 1 µm). Data in (b) and (d) is representative of 3 and 2 independent experiments respectively. Statistics (d, f) was determined by Wilcoxon matched pairs signed rank test or student’s unpaired two-sided t test (c) with P values noted in figures.

Source data

Extended Data Fig. 3

(a) Dynamic association of LAG3, TCR and CD8 on Lag3+/+ CD8+ T cells stimulated with TCRβ Ab using sensitized emission FRETc with quantification shown (Scale bar = 5 µm). (b) The average number of LAG3 and CD4 molecules that are within 25 nm x 25 nm x 50 nm area around the TCR in Lag3+/+ CD4+ T cells, isolated as above, following stimulation with TCRβ Ab as determined by the STORM-derived coordinates. Data are representative of at least 3–5 experiments. (c) Diagram depicting a fluid planar lipid bilayer system for analysis of protein–protein interactions. In the system, fluorescently tagged (TAMRA) CD4CT or CD8CT are anchored to the lipid bilayer via his-binding phospholipids resulting in fluid, randomly distributed TAMRA molecules visualized using confocal microscopy. Protein interactions are observed following addition of a peptide (LAG3CT) resulting in phase separation and redistribution of molecules into supramolecular clusters. (d, e) Molecular distribution of CD4 (d), CD8 (e) and Lck within the IS of Lag3+/+ and Lag3–/– CD4+ or CD8+ T cells, isolated as above, as determined by STORM with quantification of the Colocalization index (Scale bar = 0.5 µm). STORM imaging data are representative of 15 data sets derived from 3 separate experiments. Statistics determined by unpaired Student’s two sided t test (a, b) and by Wilcoxon matched pairs signed rank test (d, f). P values are noted in figures.

Extended Data Fig. 4

Sequence alignment of LAG3CT with acidic residues highlighted red.

Extended Data Fig. 5

(a) LAG3CT ‘EP’ and mutant peptide sequences. (b) Diagram depicting a fluid planar lipid bilayer system for analysis of the association/disassociation of p56lck from CD4 or CD8. In the system, TAMRA-labeled p56lck is associated with CD4CT or CD8CT anchored to the lipid bilayer via his-binding phospholipids resulting in fluid, randomly distributed TAMRA molecules visualized using confocal microscopy. Dissociation of p56lck is observed following addition of a peptide (LAG3CT) resulting in a reduction of TAMRA-labeled p56lck and reduced MFI. (c) Determination of phase separation and intermolecular interactions using TAMRA-labeled membrane-tethered CD4CT with AF647-labeled LAG3CT or LAG3CT-QP mutant as photoquencher. Quantification of CD4 quenching or recovery post-photobleaching of acceptor is shown from 2 independent experiments (Scale bar = 1 µm). (d) hCD4 and hCD8 CT sequence. Statistics determined by unpaired Student’s two sided t test. P values are noted in figures.

Extended Data Fig. 6

(a) The disordered cytoplasmic tail of hLAG3 binds Zn2+ weakly, in an entropically driven thermodynamic process. ITC curves at 15°C of Zn2+ titrated into 75 μM hLAG3 peptides, in 10 mM Tris pH 7.0 buffer show that acidic residues are required for metal binding. (b) Cytoplasmic tails of hLck and hCD4 fold upon binding in the presence of Zn2+ as described in Kim PW et al., Science (2003). 1H/15N HSQC spectra of a 50 μM 1:1 complex 15N-Lck: CD4, in the presence of 1.2 molar equivalents of Zn2+ (blue) and the same sample treated with 0.25 mM EDTA (red).

Extended Data Fig. 7

(a, b) Analysis of TCR-induced signaling events in Lag3–/–CD8+ T cells, isolated as above, transduced with LAG3WT or LAG3 functional domain mutants and stimulated with TCRβ Abs. Representative super-resolution confocal images are depicted, with quantification of single cell intensity measurements collected from two independent experiments presented as percent of LAG3-deficient parental cells as control (Scale bar 2 µm). Statistics determined by unpaired Student’s two sided t test. P values are noted in figures.

Supplementary information

Supplementary Information

Supplementary Tables 1–3 and Videos 1–3.

Reporting Summary.

Supplementary Video 1

TIRFM visualizing and TCR (green) of Lag3–/– CD4+ T cells containing FAP-tagged LAG3 stimulated on a planar lipid bilayer with TCRβ Ab for 15 min. Scale bar, 5 µm.

Supplementary Video 2

TIRFM visualizing LAG3 (red) of Lag3–/– CD4+ T cells isolated as above, containing FAP-tagged LAG3 stimulated on a planar lipid bilayer with TCRβ Ab for 15 min. Scale bar, 5 µm.

Supplementary Video 3

TIRFM visualizing both LAG3 (red) and TCR (green) of Lag3–/– CD4+ T cells isolated as above, containing FAP-tagged LAG3 stimulated on a planar lipid bilayer with TCRβ Ab for 15 min. Scale bar, 5 µm.

Source data

Source Data Fig. 2

Unprocessed Western blots.

Source Data Fig. 4

Unprocessed Western blots.

Source Data Fig. 5

Unprocessed Western blots.

Source Data Fig. 6

Unprocessed Western blots.

Source Data Extended Data Fig. 1

Unprocessed Western blots.

Source Data Extended Data Fig. 2

Unprocessed Western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guy, C., Mitrea, D.M., Chou, PC. et al. LAG3 associates with TCR–CD3 complexes and suppresses signaling by driving co-receptor–Lck dissociation. Nat Immunol 23, 757–767 (2022). https://doi.org/10.1038/s41590-022-01176-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-022-01176-4

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer